
A modified nonlinear incidence rate in an SIS epidemic model was investigated. When a new disease emerged or an old one resurged, the infectivity was initially high. Subsequently, the psychological effect attenuated the infectivity. Eventually, due to the crowding effect, the infectivity reached a saturation state. The nonlinearity of the functional form of the infection incidence was modified to enhance its biological plausibility. The stability of the associated equilibria was examined, and the basic reproduction number and the critical value that governed the dynamics of the model were deduced. Bifurcation analyses were presented, encompassing backward bifurcation, saddle-node bifurcation, Bogdanov-Takens bifurcation of codimension two, and Hopf bifurcation. Numerical simulations were conducted to validate our findings.
Citation: Jianzhi Cao, Xuhong Ji, Pengmiao Hao, Peiguang Wang. Bifurcation analysis of an SIS model with a modified nonlinear incidence rate[J]. Electronic Research Archive, 2025, 33(6): 3901-3930. doi: 10.3934/era.2025173
[1] | Xin Du, Quansheng Liu, Yuanhong Bi . Bifurcation analysis of a two–dimensional p53 gene regulatory network without and with time delay. Electronic Research Archive, 2024, 32(1): 293-316. doi: 10.3934/era.2024014 |
[2] | Yichao Shao, Hengguo Yu, Chenglei Jin, Jingzhe Fang, Min Zhao . Dynamics analysis of a predator-prey model with Allee effect and harvesting effort. Electronic Research Archive, 2024, 32(10): 5682-5716. doi: 10.3934/era.2024263 |
[3] | Yuan Tian, Yang Liu, Kaibiao Sun . Complex dynamics of a predator-prey fishery model: The impact of the Allee effect and bilateral intervention. Electronic Research Archive, 2024, 32(11): 6379-6404. doi: 10.3934/era.2024297 |
[4] | Xianjun Wang, Huaguang Gu, Bo Lu . Big homoclinic orbit bifurcation underlying post-inhibitory rebound spike and a novel threshold curve of a neuron. Electronic Research Archive, 2021, 29(5): 2987-3015. doi: 10.3934/era.2021023 |
[5] | Qixiang Wen, Shenquan Liu, Bo Lu . Firing patterns and bifurcation analysis of neurons under electromagnetic induction. Electronic Research Archive, 2021, 29(5): 3205-3226. doi: 10.3934/era.2021034 |
[6] | Mengxin He, Zhong Li . Dynamic behaviors of a Leslie-Gower predator-prey model with Smith growth and constant-yield harvesting. Electronic Research Archive, 2024, 32(11): 6424-6442. doi: 10.3934/era.2024299 |
[7] | Fengrong Zhang, Ruining Chen . Spatiotemporal patterns of a delayed diffusive prey-predator model with prey-taxis. Electronic Research Archive, 2024, 32(7): 4723-4740. doi: 10.3934/era.2024215 |
[8] | Qinghua Zhu, Meng Li, Fang Han . Hopf bifurcation control of the ML neuron model with Hc bifurcation type. Electronic Research Archive, 2022, 30(2): 615-632. doi: 10.3934/era.2022032 |
[9] | Xianjun Wang, Huaguang Gu . Post inhibitory rebound spike related to nearly vertical nullcline for small homoclinic and saddle-node bifurcations. Electronic Research Archive, 2022, 30(2): 459-480. doi: 10.3934/era.2022024 |
[10] | Mengting Sui, Yanfei Du . Bifurcations, stability switches and chaos in a diffusive predator-prey model with fear response delay. Electronic Research Archive, 2023, 31(9): 5124-5150. doi: 10.3934/era.2023262 |
A modified nonlinear incidence rate in an SIS epidemic model was investigated. When a new disease emerged or an old one resurged, the infectivity was initially high. Subsequently, the psychological effect attenuated the infectivity. Eventually, due to the crowding effect, the infectivity reached a saturation state. The nonlinearity of the functional form of the infection incidence was modified to enhance its biological plausibility. The stability of the associated equilibria was examined, and the basic reproduction number and the critical value that governed the dynamics of the model were deduced. Bifurcation analyses were presented, encompassing backward bifurcation, saddle-node bifurcation, Bogdanov-Takens bifurcation of codimension two, and Hopf bifurcation. Numerical simulations were conducted to validate our findings.
For thousands of years, people have suffered the ravages of infectious diseases. Construction of a mathematical model for epidemic diseases and exploration of its dynamics can offer a theoretical foundation for the formulation of prevention and control strategies. [1] investigated positivity, stability, and endemic steady-state attainability for a class of epidemic models using linear algebraic tools. [2] studied the nonnegativity and local and global stability properties of the solutions of a newly proposed epidemic model. Bacterial-transmitted diseases, such as encephalitis, gonorrhea, and bacillary dysentery, have no immunity after rehabilitation and can be reinfected. In 1932, Kermack and Mckendrick [3] presented the SIS(Susceptible-Infectious-Susceptible) model for this kind of disease, and the infected person (I) became susceptible (S) again after rehabilitation.
The incidence rate of infectious diseases constitutes one of the most crucial elements in characterizing the infectious capacity of diseases within the model. In the classical epidemic model, the incidence rate was postulated as the bilinear function f(I)S=kIS, where k denotes the probability of a susceptible individual becoming infected upon contact with an infected one. This implies that as the number of infected individuals increases, the likelihood of susceptible individuals becoming infected also increases. Such a scenario might prevail when the number of infective individuals I is relatively small. However, when the number of infective individuals I becomes exceedingly large, the bilinear function kIS fails to provide a rational description of the spread of infectious diseases. This is attributable, in part, to the advancements in modern medicine and improved public awareness of protection against infectious diseases.
In 1978, Capasso and Serio [4] proposed a nonlinear saturated incidence rate f(I)S=kIS1+αI for modeling the epidemics of cholera in Bari in 1973. Here, 11+αI represents the inhibitory effect of the protective measures adopted by susceptible individuals, and kI1+αI is known as an incidence function which will approach a saturation level kα as the number of infectious individuals I increases.
In [5,6], a generalized form of the nonmonotonic incidence rate f(I)S=kIS1+βI+αI2 was taken into account. When the number I approaches infinity, the incidence function f(I)=kI1+βI+αI2 will tend to zero, implying that the psychological effect or the inhibitory effect is overly strong. As a result, this model might not be capable of appropriately describing some specific infectious diseases like influenza.
With the aim of devising a more reasonable incidence function, which increases initially when a disease breaks out, then declines due to the psychological effect, and finally attains a saturation level because of the crowding effect, Lu et al. [7] put forward a generalized nonmonotone and saturated incidence rate in the form of kI2S1+βI+αI2. Nevertheless, in their research, the parameter β is capable of taking a negative value, which lacks a realistic biological interpretation. Moreover, it might seem rather forced to endow the incidence rate with the combined characteristics of monotonicity, nonmonotonicity, and saturation properties. In reality, in 2000, P. van den Driessche [8] et al. put forward an epidemic model whose incidence rate was given by (k1I+k2I2)S. In doing so, they accounted for the dependence of the contact rate on either the fraction of infective individuals or the level of severity of the infection among the infected individuals. M. H. Alharbi [9] formulated an ODE(Ordinary Differential Equation) model by splitting the total host population into three disjoint epidemiological classes, susceptible individuals S(t), asymptomatic infectious individuals Ia(t), and symptomatic infectious individuals Is(t), and proposed several reasonable optimal control strategies for the control and the prevention of the disease.
Inspired by the aforementioned aspects, we are going to take into account the following SIS model, which comes with a modified nonlinear incidence rate.
{dSdt=b−dS−k1I+k2I21+aI2S+μI,dIdt=k1I+k2I21+aI2S−(d+ε+μ)I, | (1.1) |
where S(t) and I(t) signify the quantities of susceptible and infective individuals at time t. It is assumed that for all t≥0, both S(t)≥0 and I(t)≥0. The basic postulates in this paper are presented as follows:
1) The positive constant b represents the rate of population recruitment.
2) The positive constant d represents the natural death rate of S(t) and I(t).
3) The modified nonlinear incidence rate is k1I+k2I21+aI2S, where k1≥0,k2>0,a>0.
4) The positive constant μ stands for the recovery rate of infective individuals.
5) The positive constant ε stands for the death rate induced by the disease of I(t).
Remark 1. In light of the analysis of previous research, we propose a more reasonable incidence rate here.
f(I)S=k1I+k2I21+aI2S, |
where f(I) combines monotonicity, nonmonotonicity, and saturation. When k1=0, f(I) monotonically increases to the saturation level k2a as I approaches infinity. If k1>0, f(I) is nonmonotonic, first rising and then falling to k2a as I goes to infinity. The basic reproduction number and parameters have clear biological meanings. This incidence rate also enriches the model with more diverse dynamical behaviors.
As far as we know, this is the pioneer study to incorporate this modified nonlinear incidence rate into an SIS model. In subsequent sections, we will first analyze the qualitative features of equilibria like existence and topological types. Then, we will explore bifurcation behaviors such as transcritical, saddle-node, Hopf, and Bogdanov-Takens bifurcations. Finally, a summary and discussion will be presented.
In this segment, we conduct an analysis on the equilibria of system (1.1) as well as their corresponding topological types. For convenience, we perform the following transformation.
S=√d+ε+μk2x,I=√d+ε+μk2y,t=1d+ε+μτ, |
and the notations
Λ=bd+ε+μ√k2d+ε+μ,m=dd+ε+μ,q=k1√k2(d+ε+μ),p=ad+ε+μk2,w=μd+ε+μ. |
We transform system (1.1) (for the sake of simplicity, we continue to use t to represent τ into the form presented below).
{dxdt=Λ−mx−qy+y21+py2x+wy,dydt=qy+y21+py2x−y, | (2.1) |
where
Λ>0,0<m<1,q≥0,p>0,0<w<1. |
Consequently, the equilibrium points of system (2.1) comply with the subsequent equations.
{Λ−mx−qy+y21+py2x+wy=0,qy+y21+py2x−y=0. |
Obviously, a disease-free equilibrium E0(Λm,0) of (2.1) always exists. For the purpose of finding the endemic equilibrium, we are required to solve the subsequent equations.
x=1+py2q+y,Ay2+By+C=0, | (2.2) |
where A=mp+1−w,B=q−qw−Λ,C=m−Λq. Obviously, we have A>0. The discriminant of (2.2) is
Δ=(q−qw−Λ)2−4(mp+1−w)(m−Λq). |
From (2.2), there are at most two endemic equilibria E1(x1,y1) and E2(x2,y2) in the model (2.1), and they can merge into a unique endemic equilibrium E∗(x∗,y∗) when Δ=0, where
y1=Λ+qw−q−√Δ2(mp+1−w),x1=1+py21q+y1,y2=Λ+qw−q+√Δ2(mp+1−w),x2=1+py22q+y2,y∗=Λ+qw−q2(mp+1−w),x∗=1+py2∗q+y∗, |
and
y1+y2=Λ+qw−qmp+1−w,y1y2=m−Λqmp+1−w. |
From Δ=0, we derive p=(q−qw−Λ)2−4(1−w)(m−Λq)4m(m−Λq). Let
p∗=(q−qw−Λ)2−4(1−w)(m−Λq)4m(m−Λq), |
from condition 0<m<1,0<w<1, and it can be concluded that p∗>0 if, and only if,
Λq<m<(q−qw−Λ)24(1−w)+Λq. |
Therefore, the following lemma is derived.
Lemma 2. There always exists a disease-free equilibrium E0(Λm,0) in system (2.1). Moreover,
1) Model (2.1) has no endemic equilibria if, and only if, one of the following conditions holds
1.1) C≥0,−B2A≤0, i.e., m≥Λq and Λ≤q−qw.
1.2) p∗>0,−B2A>0,Δ<0, i.e., Λq<m<(q−qw−Λ)24(1−w)+Λq,Λ>q−qw, and p>p∗.
1.3) p∗<0 (which means p>p∗, i.e., Δ<0) and −B2A>0, i.e., m≥(q−qw−Λ)24(1−w)+Λq and Λ>q−qw.
2) There exists a unique endemic equilibrium E∗(x∗,y∗) if, and only if, one of the following conditions holds:
2.1) p∗>0,Δ=0,−B2A>0, i.e., Λq<m<(q−qw−Λ)24(1−w)+Λq,Λ>q−qw, and p=p∗.
2.2) C=0,−B2A>0, i.e., m=Λq and Λ>q−qw.
2.3) C<0, i.e., m<Λq.
3) There exist two different endemic equilibria E1(x1,y1) and E2(x2,y2) if, and only if, the following conditions hold:
Λq<m<(q−qw−Λ)24(1−w)+Λq,Λ>q−qw, and p<p∗, where 0<y1<y∗<y2.
Remark 3. [10] The basic reproduction number (R0) is the average number of secondary infections produced by an infection throughout the period of infection. System (2.1) always exists a disease-free equilibrium E0(Λm,0). Hence, R0 can be calculated using the next-generation matrix approach [11]. Let X=(y,x)T, then model (2.1) can be written as
dXdt=F(X)−V(X), |
where
F(X)=(qy+y21+py2x0),V(X)=(y−Λ+mx+qy+y21+py2x−wy). |
Furthermore, we can calculate that
F=(Λqm)1×1,V=(1)1×1. |
So, we can get
R0=ρ(FV−1)=Λqm. |
We note that Δ=0 (i.e., p=p∗) if, and only if, R0=1−(q−qw−Λ)24m(mp+1−w). We denote the critical value 1−(q−qw−Λ)24m(mp+1−w) by Ra, and it can be obtained that 0<Ra<1. Moreover, Ra>0 is equivalent to p>p0, where
p0=(q−qw−Λ)2−4m(1−w)4m2. |
Note that p>p0, i.e., k1>k∗1, where
k∗1=bk2d+ε−1d+ε√(d+ε)k2+ad(d+ε+μ)d. |
Hence, Lemma 2 can be written as the following result:
Theorem 4. With regard to model (2.1), the existence of a disease-free equilibrium E0(Λm,0) is always ensured. Moreover,
1) there are no endemic equilibria if one of the following conditions holds: R0≤1, Λ≤q−qw or R0<1, Λ>q−qw,R0<Ra;
2) there exists a unique endemic equilibrium E∗ if one of the following conditions holds: R0=1, Λ>q−qw, or R0>1;
3) there exist two endemic equilibria E1 and E2 if R0<1,Λ>q−qw,R0>Ra.
Remark 5. Based on Theorem 4 and system (1.1), it should be noted that the inequality Λ≤q−qw is equivalent to k1≥k10, where k10=bk2d+ε. This implies that when the ratio k1k2, which represents the proportion of linear over nonlinear infections, is greater than or equal to bd+ε and R0≤1 (meaning that, on average, an infected individual generates less than or equal to one new infected individual throughout its infectious period), the infection will not spread, that is, the disease cannot infiltrate the population.
We will explore the topological type as well as the global properties of the disease-free equilibrium.
Theorem 6. For system (2.1), the equilibrium E0(Λm,0) is
R0<1: an attracting node;
R0>1: a hyperbolic saddle point;
R0=1 and Λ≠q−qw: a saddle node of com-dimension 1;
R0=1 and Λ=q−qw: a repelling node of com-dimension 2.
Proof. The Jacobian matrix of the system (2.1) at equilibrium E(x,y) is given by
J(E)=[−m−qy+y21+py2−(q+2y)(1+py2)−(qy+y2)2py(1+py2)2x+wqy+y21+py2(q+2y)(1+py2)−(qy+y2)2py(1+py2)2x−1]. | (2.3) |
By substituting x with Λm and y with 0 in (2.3), respectively, we obtain
J(E0)=[−m−Λqm+w0Λqm−1]. |
Matrix J(E0) has two eigenvalues λ1=−m and λ2=Λqm−1=R0−1. Then, if R0<1, the disease-free equilibrium E0 is an attracting node. If R0>1, disease-free equilibrium E0 is a hyperbolic saddle point. If R0=1, the second eigenvalue is zero. To determine the type of E0, first, we linearize the system (2.1) at E0, let u=x−Λm,v=y and, using Taylor expansion, the following system is obtained.
{dudt=−mu+(w−1)v−quv−Λmv2−uv2+pv3+o(|u,v|4),dvdt=quv+Λmv2+uv2−pv3+o(|u,v|4). | (2.4) |
Next, we diagonalize the linear part of the system (2.4), let X=v,Y=u−w−1mv, and system (2.4) is transformed into
{dXdt=Λ+qw−qmX2+qXY+(w−1m−p)X3+X2Y+o(|X,Y|4),dYdt=−mY−w+m−1m(Λ+qw−qmX2+qXY+(w−1m−p)X3+X2Y)+o(|X,Y|4). | (2.5) |
Let τ=−mt (for simplicity, we still use t for τ), and we change system (2.5) to the following form:
{dXdt=−Λ+qw−qm2X2−qmXY−w−1−mpm2X3−1mX2Y+o(|X,Y|4),dYdt=Y+w+m−1m2(Λ+qw−qmX2+qXY+w−1−mpmX3+X2Y)+o(|X,Y|4). |
The system satisfies the condition for the existence of a local center manifold. Assuming that the center manifold of the system is Y=h(X), it can be obtained that
dXdt=−Λ+qw−qm2X2+(q(w+m−1)(Λ+qw−q)m4−w−1−mpm2)X3+(q(w+m−1)(w−1−mp)m4+(w+m−1)(Λ+qw−q)m4)X4+(w+m−1)(w−1−mp)m4X5+o(X6). | (2.6) |
Hence, E0 is a saddle node of com-dimension 1 when Λ≠q−qw [12].
If R0=1 and Λ=q−qw, model (2.6) becomes
dXdt=−w−1−mpm2X3+o(X4). |
Obviously, the coefficient is −w−1−qpΛm2>0; therefore, according to Theorem 7.1 in [13], E0 is a repelling node of com-dimension 2.
Theorem 7. For the system (2.1), if R0<1, the disease-free equilibrium E0 is globally asymptotically stable, that is, the disease dies out.
Proof. Set N(t)=x(t)+y(t) and sum up the first equation and the second equation of system (2.1). Then, we obtain
dNdt+mN=Λ+(m+w−1)y, |
We know that m+w−1<0, which implies
dNdt+mN≤Λ. |
By the theory of ODEs, we have
N(t)≤N(0)e−mt+Λm(1−e−mt)≤N(0)+Λm, |
where N(0) is the initial value of N(t) at t=0. Therefore, the positive invariant set of system (2.1) is
Ω={(x,y)|x(t)+y(t)≤N(0)+Λm,x(t)≥0,y(t)≥0}. |
In the case where R0<1, as per Theorem 4, it is known that the system lacks an endemic equilibrium. Furthermore, according to Theorem 6, when R0<1, the disease-free equilibrium E0 is an attracting node. Given that the disease-free equilibrium E0 is located on the boundary of Ω, it can be deduced from the Poincarˊe-Bendixson theorem that for every positive solution of system (2.1), it will approach the equilibrium E0 as t tends to +∞.
Remark 8. It can be observed that by choosing the parameters Λ=0.3, m=0.3, q=0.5, w=0.6, and p=1, the condition of Theorem 7 is fulfilled. In other words, when R0<1, the disease-free equilibrium E0 is globally asymptotically stable, which means the disease will eventually die out. The simulation of this phenomenon is presented in Figure 1.
Next, we proceed to investigate the topological type of the endemic equilibrium. As we can learn from (2.3), we are aware of the Jacobian matrix of system (2.1) at the equilibrium E(x,y). Subsequently, the determinant of J(E) is
det(J(E))=(2mp+2−2w)y2+(q−qw−Λ)y1+py2, |
and its sign is determined by
SD=(2mp+2−2w)y2+(q−qw−Λ)y. | (2.7) |
Similarly, we obtain the trace of J(E),
tr(J(E))=(−m2p−m−2pm+w−1)y2+(Λ−mq)y−m2m(1+py2), |
and its sign is determined by
ST=(−m2p−m−2pm+w−1)y2+(Λ−mq)y−m2. | (2.8) |
Theorem 9. For system (2.1), according to Lemma 2, if Λq<m<(q−qw−Λ)24(1−w)+Λq,Λ>q−qw, and p<p∗ holds, there exist two endemic equilibria E1(x1,y1) and E2(x2,y2). Furthermore, E1(x1,y1) must be a hyperbolic saddle, and then E2(x2,y2) will be
1) a stable hyperbolic focus (or node) if trJ(E2)<0; or
2) a weak focus (or a center) if trJ(E2)=0; or
3) an unstable hyperbolic focus (or node) if trJ(E2)>0.
Proof. Upon computing the determinants of J(E1) and J(E2), it turns out that their signs are determined by SD in (2.7). Through several straightforward calculations, we derive
SD(y1)=(2mp+2−2w)y21+(q−qw−Λ)y1<0,SD(y2)=(2mp+2−2w)y22+(q−qw−Λ)y2>0. |
Therefore, the endemic equilibrium E1 is a hyperbolic saddle.
(i) If trJ(E2)<0, the eigenvalues of J(E2) have negative real parts, so E2 is a stable hyperbolic focus (or node);
(ii) If trJ(E2)=0, the eigenvalues of J(E2) are a pair of pure imaginary roots, so E2 is a weak focus (or a center);
(iii) If trJ(E2)>0, the eigenvalues of J(E2) have positive real parts, so E2 is an unstable hyperbolic focus (or node).
Remark 10. When m=Λq and Λ>q−qw, or m<Λq, system (2.1) has a unique positive equilibrium E2(x2,y2) along with a disease-free equilibrium E0. E0 is either a hyperbolic saddle or saddle-node. The nature and stability of E2 accord with those in Theorem 9. Furthermore, by Theorem 9, if R0>1 (i.e., m<Λq), signifying that an infected individual generates more than one new infection on average, or if R0=1 and k1k2<bd+ε, implying that the proportion of linear to nonlinear infection hazards is less than bd+ε and an infected individual yields one new infection on average, the disease will endure as multiple periodic coexisting oscillations bifurcating from E2.
The topological types of the endemic equilibrium E∗(x∗,y∗) will be discussed in the following section.
In this section, we will conduct a bifurcation analysis of system (2.1).
Theorem 11. For the system (2.1), we choose R0 as the bifurcation parameter. When R0=1,
1) system (2.1) undergoes forward bifurcation if Λ<q−qw, and undergoes backward bifurcation if Λ>q−qw;
2) system (2.1) undergoes pitchfork bifurcation if Λ=q−qw.
Proof. Given that R0 can be regarded as a function depending on the parameters Λ, q, and m, without sacrificing generality, we may select m as the bifurcation parameter. Set m=Λq+δ, where the case of δ=0 corresponds to R0=1.
We linearize the system (2.1) at E0. First, let u′=x−ΛΛq+δ,v′=y (for simplicity, we still use u,v for u′,v′), and using Taylor expansion, the following system is obtained.
{dudt=a11u+a12v+a13uv+a14v2+a15uv2+a16v3+o(|u,v|4),dvdt=a21v+a22uv+a23v2+a24uv2+a25v3+o(|u,v|4), | (3.1) |
where
a11=−(Λq+δ),a12=w−ΛqΛq+δ,a13=−q,a14=−ΛΛq+δ,a15=−1,a16=ΛqpΛq+δ,a21=−δΛq+δ,a22=q,a23=ΛΛq+δ,a24=1,a25=−ΛqpΛq+δ. |
Next, we diagonalize the linear part of the system (3.1), let X=v,Y=u−w(Λq+δ)−Λq(Λq+δ)2−δv, τ=−(Λq+δ)t (for simplicity, we still use t for τ), and the system (3.1) is transformed into
{dXdt=b11X+b12XY+b13X2+b14X2Y+b15X3+o(|X,Y|4),dYdt=b21Y+b22XY+b23X2+b24X2Y+b25X3+o(|X,Y|4), |
where
b11=δ(Λq+δ)2,b12=−qΛq+δ,b13=−1Λq+δ(wq(Λq+δ)−Λq2(Λq+δ)2−δ+ΛΛq+δ),b14=−1Λq+δ,b15=−1Λq+δ(wq(Λq+δ)−Λq2(Λq+δ)2−δ−ΛqpΛq+δ),b21=1,b22=qΛq+δ(wq(Λq+δ)−Λq2(Λq+δ)2−δ+1),b23=1Λq+δ(wq(Λq+δ)−Λq2(Λq+δ)2−δ+1)(wq(Λq+δ)−Λq2(Λq+δ)2−δ+ΛΛq+δ),b24=1Λq+δ(wq(Λq+δ)−Λq2(Λq+δ)2−δ+1),b25=1Λq+δ(wq(Λq+δ)−Λq2(Λq+δ)2−δ+1)(wq(Λq+δ)−Λq2(Λq+δ)2−δ−ΛqpΛq+δ). |
The following reduced model on the center manifold can be obtained by applying the center manifold theorem with the parameter δ,
dXdt=δ(Λq+δ)2X−1Λq+δ(wq(Λq+δ)−Λq2(Λq+δ)2−δ+ΛΛq+δ)X2+o(X3). | (3.2) |
Denoting the righthand side of the model (3.2) as F(X,δ), we can derive
F(0,0)=0,∂F∂X(0,0)=0,∂F∂δ(0,0)=0,∂2F∂X∂δ(0,0)=1(Λq)2≠0,∂2F∂2X(0,0)=−2Λ+qw−q(Λq)2. |
Therefore, model (3.2) undergoes a transcritical bifurcation([14]) if Λ≠q−qw.
Since ∂R0∂ε=−1Λq<0, when R0 crosses R0=1, the system (2.1) undergoes a forward bifurcation if Λ<q−qw, and a backward bifurcation if Λ>q−qw, respectively.
If Λ=q−qw, model (3.2) in the center manifold becomes
dXdt=b31X+b32X2+b33X3+o(X4), | (3.3) |
where
b31=δ(q2−q2w+δ)2,b32=−q3δ−q3wδ+qδ2−qδ−qwδ((q2−q2w+δ)2−δ)(q2−q2w+δ)2,b33=−2wq3−w2q3+wqδ−q3(q2−q2w+δ)3−δ(q2−q2w+δ)−q2p−q2wp(q2−q2w+δ)2+q4δ−q4wδ+q2δ2−q2δ−q2wδ((q2−q2w+δ)2−δ)(q2−q2w+δ)3(2wq3−w2q3+wqδ−q3(q2−q2w+δ)2−δ+1). |
For simplicity, we still denote the right side of the model (3.3) as F(X,ε), and derive
F(0,0)=0,∂F∂X(0,0)=0,∂F∂δ(0,0)=0,∂2F∂X∂δ(0,0)=1(q2−q2w)2≠0,∂2F∂2X(0,0)=0,∂3F∂3X(0,0)=qp+1q3(1−w)≠0. |
Therefore, model (3.3) undergoes pitchfork bifurcation([14]) if Λ=q−qw.
Remark 12. In epidemiological models, the backward bifurcation is an important phenomenon, and the value of R0≤1 is not sufficient to reflect the prevalence of the disease. Theorem 11 proves the conditions for the occurrence of a backward bifurcation. The existence of backward bifurcation indicates that even if some parameters control the basic reproduction number R0<1, the disease will still spread and certain parameters need to be controlled to make R0<Ra to eliminate the disease.
Theorem 13. For system (2.1), according to Lemma 2, if Λq<m<(q−qw−Λ)24(1−w)+Λq,Λ>q−qw, and p=p∗ hold, there is a unique endemic equilibrium E∗(x∗,y∗). Furthermore,
1) if w≠w∗, then E∗(x∗,y∗) is a saddle node, which is attracting (or repelling) if w<w2, or w>w∗ (or w2<w<w∗);
2) if w=w∗, then E∗(x∗,y∗) is a cusp of codimension two.
Proof. We plug y=y∗,p=p∗ in (2.7) and (2.8), and then obtain SD(y∗)=0 and
ST(y∗)=C1w2+C2w+C3m(Λ+qw−q)2, |
where
C1=q2(−2m3−2m2+m2Λq+2mΛq),C2=2q(Λ−q)(−2m3−2m2+m2Λq+2mΛq)−(4m+4m2)(m−Λq)2+2mq(Λ−qm)(m−Λq),C3=(Λ−q)2(−2m3−2m2+m2Λq+2mΛq)+4m(m−Λq)2+2m(Λ−q)(Λ−qm)(m−Λq). |
Therefore, the symbol of ST(y∗) is determined by C1w2+C2w+C3, and we denote
F(w)=C1w2+C2w+C3. |
By calculation,
ΔF(w)=4m2(m−Λq)2(2m(m+1)−q(Λ+mq))2>0. |
It is easy to determine that 2m(m+1)−q(Λ+mq)>0, then
√ΔF(w)=2m(m−Λq)(2m(m+1)−q(Λ+mq)). |
We assume that the two roots of F(w) are w1 and w2, and we can get
w1=−C2−√ΔF(w)2C1=2m−2Λq−mΛ2+mΛq2m2+2m−2Λq−mΛq,w2=−C2+√ΔF(w)2C1=q2+Λq−2mq2. |
Since Λq<m<(q−qw−Λ)24(1−w)+Λq,Λ>q−qw, we can get 0<w2<w1<1; moreover, when w=w2, it conflicts with the condition m>Λq. Hence, we denote
w∗=2m−2Λq−mΛ2+mΛq2m2+2m−2Λq−mΛq. |
Therefore, if w=w∗,ST(y∗)=0, ST(y∗)<0 if, and only if, w<w2 or w>w∗. Conversely, if w2<w<w∗,ST(y∗)>0. Using the transformation of u1=y−y∗,u2=x−x∗,p=p∗, we rewrite system (2.1) as follows:
{du1dt=q(u1+y∗)+(u1+y∗)21+p∗(u1+y∗)2(u2+x∗)−(u1+y∗),du2dt=Λ−m(u2+x∗)−q(u1+y∗)+(u1+y∗)21+p∗(u1+y∗)2(u2+x∗)+w(u1+y∗). | (3.4) |
Letting U1=u1,U2=u1+u2 (U1,U2 are rewritten as u1,u2) and using Taylor expansion, then the following system is obtained.
{du1dt=c11u1+c12u2+c13u21+c14u1u2+o(|u1,u2|3),du2dt=c21u1+c22u2+o(|u1,u2|3), | (3.5) |
where
c11=−4(m−Λq)2(m+w−1)−2q(m−Λq)(Λ+qw−q)(m+w−1)(2m−Λq)(Λ+qw−q)2−4(1−w)(m−Λq)2,c12=2mq(m−Λq)(Λ+qw−q)+4m(m−Λq)2(2m−Λq)(Λ+qw−q)2−4(1−w)(m−Λq)2,c13=−3(m−Λq)3−2(Λ−mq)(Λ+qw−q)2+8(Λ+qw−q)(m−Λq)(1−m−w)(2m−Λq)(Λ+qw−q)2−4(1−w)(m−Λq)2+−4m(Λ+qw−q)5(m−Λq)+4m(mq−Λ)(Λ+qw−q)4(m−Λq)m((2m−Λq)(Λ+qw−q)2−4(1−w)(m−Λq)2)2+(40m−40mw+8m2)(Λ+qw−q)3(m−Λq)2m((2m−Λq)(Λ+qw−q)2−4(1−w)(m−Λq)2)2−32m(1−w)(3−3w+m)(Λ+qw−q)(m−Λq)3m((2m−Λq)(Λ+qw−q)2−4(1−w)(m−Λq)2)2,c14=mΛq2(Λ+qw−q)4+4m2(m−Λq)(Λ+qw−q)3((2m−Λq)(Λ+qw−q)2−4(1−w)(m−Λq)2)2+4mq(1−w)(m−Λq)2(Λ+qw−q)2((2m−Λq)(Λ+qw−q)2−4(1−w)(m−Λq)2)2,c21=m+w−1,c22=−m. |
Defining
κ1=4(1−w)(m−Λq)2−(2m−Λq)(Λ+qw−q)24(m−Λq)2+2q(m−Λq)(Λ+qw−q),κ2=m+w−1m, |
and letting
ρ1=κ1u1−u2κ1−κ2,ρ2=−κ2u1+u2κ1−κ2. |
We transform (3.5) into the following form:
{dρ1dt=d11ρ21+d12ρ1ρ2+d13ρ22+o(|ρ1,ρ2|3),dρ2dt=d20ρ2+d21ρ21+d22ρ1ρ2+d23ρ22+o(|ρ1,ρ2|3), | (3.6) |
where
d11=κ1(κ2c14+c13)κ1−κ2,d12=κ1((κ1+κ2)c14+2c13)κ1−κ2,d13=κ1(κ1c14+c13)κ1−κ2,d20=−κ1κ2c12+κ1c22−κ2c11+c21κ1−κ2,d21=−κ2(κ2c14+c13)κ1−κ2,d22=−κ2((κ1+κ2)c14+2c13)κ1−κ2,d23=−κ2(κ1c14+c13)κ1−κ2. |
We further compute d20 and obtain
d20=(m+w−1)(4(1−w)(m−Λq)2−(2m−Λq)(Λ+qw−q)2)4(m−Λq)2+2q(m−Λq)(Λ+qw−q)−m. |
When w≠w∗, d20≠0. Introduce a new time variable τ through the relation dτ=d20dt, and then rewrite τ as t. From (3.6), we obtain
{dρ1dt=e11ρ21+e12ρ1ρ2+e13ρ22+o(|ρ1,ρ2|3),dρ2dt=ρ2+e21ρ21+e22ρ1ρ2+e23ρ22+o(|ρ1,ρ2|3), | (3.7) |
where eij=dijd20 (i=1,2;j=1,2,3). Letting the right side of the second equation system (3.7) be equal to zero, we can solve the implicit function ρ2=Φ(ρ1) as follows:
ρ2=Φ(ρ1)=−e21ρ21+⋅⋅⋅. | (3.8) |
Substituting (3.8) into the right side of the first equation of the system (3.7), we get
dρ1dt=e11ρ21+e12ρ1Φ(ρ1)+e13Φ2(ρ1)+⋅⋅⋅, |
where
e11=κ1(κ2c14+c13)−κ1κ2c12+κ1c22−κ2c11+c21. |
After calculation, it can be concluded that if w≠w∗, e11≠0. Therefore, according to Theorems 7.1–7.3 in Zhang [13] et al., E∗ is a saddle node of codimension one. Then, we obtain the conclusion in (1).
In addition, the saddle-node bifurcation of system (2.1) will occur as the parameter p crosses the critical value p∗ from the right to the left.
For the second conclusion in (2), substituting w=w∗ into (3.4) and applying the Taylor expansion, the system (3.4) is transformed into
{du1dt=a1u1+a2u2+a3u21+a4u1u2+o(|u1,u2|3),du2dt=−a21a2u1−a1u2−a3u21−a4u1u2+o(|u1,u2|3), | (3.9) |
where a1,a2,a3,a4 are listed in the Appendix.
Letting v1=u1,v2=a1u1+a2u2, the system (3.9) is transformed into
{dv1dt=v2+b1v21+b2v1v2+o(|v1,v2|3),dv2dt=b3v21+b4v1v2+o(|v1,v2|3), | (3.10) |
where
b1=a2a3−a1a4a2,b2=a4a2,b3=(a1−a2)(a2a3−a1a4)a2,b4=(a1−a2)a4a2, |
and the parameter a1,a2,a3,a4 is the same as in system (3.9). To eliminate the quadratic terms in the first equation of the system (3.10), we transform η1=v1−b22v21,η2=v2+b1v21 and obtain
{dη1dt=η2+o(|η1,η2|3),dη2dt=b3η21−(2b1+b4)η1η2+o(|η1,η2|3). |
From conditions w=w∗,Λq<m<(q−qw−Λ)24(1−w)+Λq,Λ>q−qw, we have
b3≠0,2b1+b4≠0. |
Due to the complexity of the formula, the calculation process is omitted here. From the results in [15], E∗(x∗,y∗) is a cusp of codimension two.
Next, we select parameter values of Λ,m,q,w, and p to simulate the result of Theorem 13. First, letting Λ=0.2,m=0.1,q=0.45,p=4.129, and w=0.88, we find that E∗ is an attracting saddle node, and the phase portrait is shown in Figure 2. Second, letting Λ=0.2,m=0.1,q=0.4,p=0.712, and w=0.84, the phase portrait of the repelling saddle-node is shown in Figure 3. Finally, letting Λ=0.4,m=0.2,q=0.4,p=564, and w=0.625, we can obtain E∗ as a codimension two cusp, and the phase portrait is shown in Figure 4.
Remark 14. When q=0 (that is, k1=0), system (2.1) is similar to model (1.3) of Tang et al. [16]. From Theorem 13, we prove that model (2.1) undergoes the Bogdanov-Takens bifurcation of codimension at most two near E∗(x∗,y∗). From Theorems 11 and 13, when k∗1<k1<k10 (i.e., Λ>q−qw) for system (2.1), the disease will be eliminated if R0<Ra (i.e., p>p∗ and k1>k∗1), and if R0=Ra (i.e., p=p∗ and k1>k∗1), system (2.1) will present complex dynamics; these conditions are not enough to determine dynamical behaviors and the disease will persist or die out, depending on the values of k1 and k2.
In this section, we discuss Hopf bifurcation at endemic equilibrium E2(x2,y2). First, we introduce the new state variable u=x+y−Λm,v=y, time transformation dt=(1+py2)dτ, and change the system (2.1) into the following form:
{dudt=−mu+(m+w−1)v−mpuv2+(m+w−1)pv3,dvdt=(Λqm−1)v+(Λm−q)v2+quv+uv2−(1+p)v3. | (3.11) |
Since 1+py2>0, system (3.11) is topologically equivalent to the system (2.1) and has an endemic equilibrium ˜E2(u2,v2) corresponding to E2(x2,y2) of (2.1), where
u2=m+w−1mv2,v2=Λ+qw−q+√(q−qw−Λ)2−4(mp+1−w)(m−Λq)2(mp+1−w). |
The Jacobian matrix of the system (3.11) at equilibrium ˜E2(u2,v2) is given by
J(˜E2)=[−m−mpv22(m+w−1)(1+pv22)qv2+v22Λq−mm+2Λ+wq−mq−qmv2+−3mp−m+2w−2mv22] |
We know that
sgn(detJ(˜E2))=sgn((1+pv22)((Λ+qw−q)v2−2(m−Λq))=sgn((q−qw−Λ)2−4(mp+1−w)(m−Λq)+√(q−qw−Λ)2−4(mp+1−w)(m−Λq))=1,sgn(trJ(˜E2))=−sgn((mp+m+m2p)v22+(mq−q+qw)v2+Λq−m+m2), |
and sgn(trJ(˜E2))=0 is equivalent to
(mp+m+m2p)v22+(mq−q+qw)v2+Λq−m+m2=0. | (3.12) |
Since Λq<m<(q−qw−Λ)24(1−w)+Λq,Λ>q−qw, we can obtain Λq−m+m2<0; thus, (mq−q+qw)2−4(mp+m+m2p)(Λq−m+m2)>0. Eq (3.12) has two real solutions
v2±=−(mq−q+qw)±√(mq−q+qw)2−4(mp+m+m2p)(Λq−m+m2)2(mp+m+m2p), |
where v+2>0,v−2<0. Since v2>0, in determining the sign of trJ(˜E2), we may consider the relationship between v2 and v+2 as follows:
v2>v+2,v2=v+2,v2<v+2, |
which are equivalent to ϑ>ϑ∗,ϑ=ϑ∗,ϑ<ϑ∗, respectively, where
ϑ≜(mp+m+m2p)(Λ+qw−q+√(q−qw−Λ)2−4(mp+1−w)(m−Λq)),ϑ∗≜(mp+1−w)(−(mq−q+qw)+√(mq−q+qw)2−4(mp+m+m2p)(Λq−m+m2)). |
Therefore, we can come to the following conclusion: if ϑ>ϑ∗, then sgn(trJ(˜E2))<0; if ϑ=ϑ∗, then sgn(trJ(˜E2))=0; if ϑ<ϑ∗, then sgn(trJ(˜E2))>0. Then, we have the following theorem.
Theorem 15. Equilibrium ˜E2(x2,y2) satisfies one of the following cases:
1) when ϑ>ϑ∗, equilibrium ˜E2 is a locally stable focus or node;
2) when ϑ=ϑ∗, equilibrium ˜E2 is a weak focus or center;
3) when ϑ<ϑ∗, equilibrium ˜E2 is a locally unstable focus or node.
Next, we discuss the Hopf bifurcation around the equilibrium ˜E2. From expressions of tr(˜E2) and ϑ, we can represent tr(˜E2) as a function of ϑ,
trJ(˜E2)=−(mp+m+m2p)(ϑ2(mp+m+m2p)(mp+1−w))2−(mq−q+qw)(ϑ2(mp+m+m2p)(mp+1−w))−(Λq−m+m2). |
We compute the derivative of trJ(˜E2) with respect to ϑ at ϑ∗ and obtain
trJ(˜E2)dϑ|ϑ=ϑ∗=−√(mq−q+qw)2−4(mp+m+m2p)(Λq−m+m2)2(mp+m+m2p)(mp+1−w)<0. |
Using a transformation of φ=u−u2,ψ=v−v2 and Taylor expansion, system (3.11) can be changed into
{dφdt=a11φ+a12ψ+a13φψ+a14ψ2+a15φψ2+a16ψ3,dψdt=a21φ+a22ψ+a23φψ+a24ψ2+a25φψ2+a26ψ3, | (3.13) |
where
a11=−m−mpv22,a12=(m+w−1)(1+pv22),a13=−2mpv2,a14=2p(m+w−1)v2,a15=−mp,a16=p(m+w−1),a21=qv2+v22,a22=Λq−mm+2Λ+wq−mq−qmv2+−3mp−m+2w−2mv22,a23=q+2v2,a24=Λ−mqm+−3mp−2m+w−1mv2,a25=1,a26=−(1+p). |
Defining D=√detJ(˜E2) and introducing transformation of η=−φ,ρ=a11Dφ+a12Dψ, we obtain, from system (3.13), that
{dηdt=−Dρ+f(η,ρ),dρdt=Dη+g(η,ρ), |
where
f(η,ρ)=b11ηρ+b12ρ2+b13η2ρ+b14ηρ2+b15ρ3,g(η,ρ)=b21η2+b22ηρ+b23ρ2+b24η3+b25η2ρ+b26ηρ2+b27ρ3, |
and
b11=2mpv2D(m+w−1)(1+pv22),b12=−2pv2D2(m+w−1)(1+pv22)2,b13=−m2pD(m+w−1)2(1+pv22),b14=2mpD2(m+w−1)2(1+pv22)2,b15=−pD3(m+w−1)2(1+pv22)3,b21=m(1+pv22)((−3mp+3w+3)v2+Λ+qw−q)(m+w−1)D,b22=(2m2p+6mp+2m−4w+4)v2+mq−2Λ−qw+qm+w−1,b23=((−2m2p−3mp−2m+w−1)v2+Λ−mq)Dm(m+w−1)(1+pv22),b24=m2(1+pv22)(mp−w+1)(m+w−1)2D,b25=−m(m2p+3mp+m+2−2w)(m+w−1)2,b26=(2m2p+3mp+2m−w+1)D(m+w−1)2(1+pv22),b27=−(mp+p+1)D(m+w−1)2(1+pv22)2. |
Based on the results in [17,18], the formula of the first Lyapunov number is given by
l1=(fηηη+fηρρ+fηηρ+fρρρ16+fηρ(fηη+fρρ)−gηρ(gηη+gρρ)−fηηgηη+fρρgρρ16D)|η=0,ρ=0. |
With a simple computation, we have
l1=σ8(m+w−1)2, |
where
σ=2mpD2(1+pv22)2−m2pD1+pv22−3pD3(1+pv22)3−4mp2v22D2(1+pv22)3−1D2((m(1+pv22)((−3mp+3w+3)v2+Λ+qw−q))((2m2p+6mp+2m−4w+4)v2+mq−2Λ−qw+q))−(((2m2p+6mp+2m−4w+4)v2+mq−2Λ−qw+q)m(1+pv22)×((−2m2p−3mp−2m+w−1)v2+Λ−mq))−4pv2D((−2m2p−3mp−2m+w−1)v2+Λ−mq)m(1+pv22)3. |
Based on the above discussion, we have the following theorem.
Theorem 16. Assuming Λq<m<(q−qw−Λ)24(1−w)+Λq,Λ>q−qw, and p<p∗ hold,
1) when σ>0 and the parameter ϑ passes through ϑ∗ from left to right, the system (3.11) undergoes a subcritical Hopf bifurcation around the equilibrium ˜E2 in the sufficiently small neighborhood of ϑ∗, and an unstable limit cycle appears.
2) when σ<0 and the parameter ϑ passes through ϑ∗ from right to left, the system (3.11) undergoes a supercritical Hopf bifurcation around the equilibrium ˜E2 in the sufficiently small neighborhood of ϑ∗, and a stable limit cycle appears.
We take parameters of Λ=0.2,m=0.1,q=0.45,p=1,w=0.8, then σ>0 and ϑ<ϑ∗. According to (1) of Theorem 16, the system (2.1) has an unstable focus (see Figure 5 (a)). When we change the parameter m from 0.1 to 0.0998, the relationship of ϑ<ϑ∗ becomes ϑ>ϑ∗, and the system (2.1) exists as an unstable limit cycle (see Figure 5 (b)).
We take parameters of Λ=0.2,m=0.099,q=0.45,p=1,w=0.8, then σ<0 and ϑ>ϑ∗. According to (2) of Theorem 16, (2.1) has a stable focus (see Figure 6 (a)). When we take parameters of Λ=0.5,m=0.1,q=0.1,p=0.1,w=0.517, then σ<0 and ϑ<ϑ∗, and system (2.1) exists as a stable limit cycle (see Figure 6 (b)).
From the conclusion (2) of Theorem 13 in the previous section, we know that the endemic equilibrium E∗ is a codimension two-cusp point. It indicates that system (2.1) may exhibit the Bogdanov-Takens bifurcation under a small parameter perturbation. In this section, we study whether this bifurcation phenomenon can occur in the small neighborhood of E∗ or not. Let Λ0,m0,p0,q0, and w0 be the fixed values for the parameters Λ,m,p,q, and w, respectively, to ensure that conditions p=p∗ and w=w∗ hold, that is,
p0=(q0−q0w0−Λ0)2−4(1−w0)(m0−Λ0q0)4m0(m0−Λ0q0),w0=2m0−2Λ0q0−m0Λ20+m0Λ0q02m20+2m0−2Λ0q0−m0Λ0q0. |
Selecting Λ and m as bifurcation parameters, θ1 and θ2 as their perturbations at Λ0 and m0, respectively, and using the transformation of u1=y−y∗,u2=x−x∗, we have, from system (2.1), that
{du1dt=q0(u1+y∗)+(u1+y∗)21+p0(u1+y∗)2(u2+x∗)−(u1+y∗),du2dt=(Λ0+θ1)−(m0+θ2)(u2+x∗)−q0(u1+y∗)+(u1+y∗)21+p0(u1+y∗)2(u2+x∗)+w0(u1+y∗). | (3.14) |
By performing Taylor expansion, system (3.14) becomes
{du1dt=c11u1+c12u2+c13u21+c14u1u2+R1(θ1,θ2,u1,u2),du2dt=c20+c21u1+c22u2+R2(θ1,θ2,u1,u2), | (3.15) |
where R1(θ1,θ2,u1,u2) and R2(θ1,θ2,u1,u2) are C∞ functions at least of the third order concerning (u1,u2), and their coefficients smoothly depend on θ1,θ2, and cij (i=1,2;j=0,1,...,4), which are listed in the Appendix. Let
v1=u1,v2=c11u1+c12u2+c13u21+c14u1u2+R1(θ1,θ2,u1,u2), |
and system (3.15) becomes
{dv1dt=v2,dv2dt=e0+e1v1+e2v2+e3v21+e4v1v2+R3(θ1,θ2,v1,v2), | (3.16) |
where R3(θ1,θ2,v1,v2) are C∞ functions at least of the third order concerning (v1,v2), and its coefficients smoothly depend on θ1,θ2, and
e0=c12c20,e1=c14c20+c12c21−c11c22,e2=c11+c22,e3=c14c21−c22c13−c13c12+c11c14,e4=−c14+2c13−c14c11c12c212,e5=c14c12. |
Introducing a new time variable τ by dt=(1−c14c12v1)dτ and rewriting τ as t, we change the system (3.16) into
{dv1dt=(1−c14c12v1)v2,dv2dt=(1−c14c12v1)(e0+e1v1+e2v2+e3v21+e4v1v2+R3(θ1,θ2,v1,v2)). | (3.17) |
Let η1=v1,η2=(1−c14c12v1)v2, and the system (3.17) becomes
{dη1dt=η2,dη2dt=e0+γ1η1+e2η2+γ3η21+γ4η1η2+R4(θ1,θ2,η1,η2)), |
where R4(θ1,θ2,η1,η2) are C∞ functions at least of the third order concerning (η1,η2), and its coefficients smoothly depend on θ1,θ2, and
γ1=e1−2e0e5,γ3=e3−2e1e5+e0e25,γ4=e4−e2e5. |
When θ1=0,θ2=0,
γ3=(c12−c11)(c11c14−c12c13)c12,γ4=c211c14−c212c14+2c212c13c212, |
where c11,c12,c13, and c14 are the same as in the system (3.15). Substituting their values for the calculation yields γ1=0,γ2=0,γ3≠0, and γ4≠0. Due to the expression being too long, the calculation process and results are ignored here. Therefore, γ3 and γ4 are not equal to zero in the small enough neighborhood of (θ1,θ2)=(0,0). Letting ζ1=η1−γ12γ3 and ζ2=η2, we have
{dζ1dt=ζ2,dζ2dt=γ0+γ2ζ2+γ3ζ21+γ4ζ1ζ2+R5(θ1,θ2,ζ1,ζ2)), | (3.18) |
where R5(θ1,θ2,ζ1,ζ2) are C∞ functions at least of the third order concerning (ζ1,ζ2) and its coefficients depend smoothly on θ1,θ2, and
γ0=e0−γ214γ3,γ2=e2−γ4γ12γ3. |
Using the following transformations of state variables and time variables as
ρ1=γ24γ3ζ1,ρ2=γ24γ23ζ2,τ=γ3γ4t, |
and still denoting τ as t, system (3.18) becomes
{dρ1dt=ρ2,dρ2dt=μ1+μ2ρ2+ρ21+ρ1ρ2+R6(θ1,θ2,ρ1,ρ2)), | (3.19) |
where
μ1=γ0γ44γ33,μ2=γ2γ4γ3. |
From the above calculation of the coefficient parameters, μ1 and μ2 can be expressed in terms of θ1 and θ2 as follows:
{μ1=κ11θ1+κ12θ2+κ13θ21+κ14θ1θ2+κ15θ22+K1(θ1,θ2),μ2=κ21θ1+κ22θ2+κ23θ21+κ24θ1θ2+κ25θ22+K2(θ1,θ2), |
where
κ11=m0γ44c12γ33,κ12=γ44c12γ33(2m0−Λ0q0−Λ20+Λ0q0w0−2m0w0Λ0+q0w0−q0),κ13=−m20γ44c2148γ53(24γ3(c14+c12)+c12),κ14=m0γ44c148γ53(c12(24γ3−1)−c142m0−Λ0q0−Λ20+Λ0q0w0−2m0w0Λ0+q0w0−q0),κ15=−γ448γ53(c12+c142m0−Λ0q0−Λ20+Λ0q0w0−2m0w0Λ0+q0w0−q0)2,κ21=m0γ24c142γ33,κ22=γ242γ33(c12+c142m0−Λ0q0−Λ20+Λ0q0w0−2m0w0Λ0+q0w0−q0),κ23=m20γ44c314(2γ3−1)4γ43c12,κ24=m0γ24c2144γ43c12(c12(3+2γ3)+c14(2γ3−1)2m0−Λ0q0−Λ20+Λ0q0w0−2m0w0Λ0+q0w0−q0),κ25=γ24c14(2γ3−1)4γ43c12(c214(2m0−Λ0q0−Λ20+Λ0q0w0−2m0w0Λ0+q0w0−q0)2−c212). |
By calculation, we get
|∂(μ1,μ2)∂(θ1,θ2)|(θ1,θ2)=(0,0)=m0γ64c2122γ63≠0. |
It shows that parameters μ1 and μ2 are independent and reversible in a small neighborhood of the origin of (θ1,θ2). By the results in Bogdanov [19] and Takens [20], we can see that the system (3.19) (i.e., (2.1)) undergoes the Bogdanov-Takens bifurcation when (θ1,θ2) changes in a small neighborhood of (0,0). That is the following theorem.
Theorem 17. When Λq<m<(q−qw−Λ)24(1−w)+Λq,Λ>q−qw,p=p∗, and w=w∗ hold, the system (2.1) has a cusp E∗(x∗,y∗) of codimension two (i.e., Bogdanov-Takens singularity). Thus, system (2.1) undergoes the Bogdanov-Takens bifurcation of codimension two around the endemic equilibrium E∗. Specifically, if Λ and m are selected as the bifurcation parameters, the system has an unstable limit cycle for some parameter values of Λ and m, and it has an unstable homoclinic loop for some other parameter values of Λ and m.
Based on the study of Perko [21], the approximate representations up to the second order of the bifurcation curves are given by the following.
(ⅰ) The saddle-node bifurcation curve is:
SN={(μ1,μ2)|μ1=0,μ2≠0}={(θ1,θ2)|κ11θ1+κ12θ2+κ13θ21+κ14θ1θ2+κ15θ22=0,μ2≠0}, |
(ⅱ) The Hopf bifurcation curve is:
H={(μ1,μ2)|μ2=√−μ1,μ1<0}={(θ1,θ2)|κ21θ1+κ22θ2+κ23θ21+κ24θ1θ2+κ25θ22−√−κ11θ1−κ12θ2−κ13θ21−κ14θ1θ2−κ15θ22=0,μ1<0}, |
(ⅲ) The homoclinic bifurcation curve is:
HL={(μ1,μ2)|μ2=57√−μ1,μ1<0}={(θ1,θ2)|κ21θ1+κ22θ2+κ23θ21+κ24θ1θ2+κ25θ22−57√−κ11θ1−κ12θ2−κ13θ21−κ14θ1θ2−κ15θ22=0,μ1<0}. |
The bifurcation diagram and the corresponding phase portraits of the system (2.1) are presented below.
(ⅰ) In Figure 7, the small neighborhood of origin in the parameter (θ1,θ2)-plane is divided into four regions by bifurcation curves of SN(Saddle-Node), H(Hopf), and HL(Homoclinic).
(ⅱ) When (θ1,θ2)=(0,0), Figure 4 shows the system (2.1) having a unique endemic equilibrium point, which is a cusp of codimension two.
(ⅲ) If the perturbation parameters (θ1,θ2) fall in area Ⅰ, system (2.1) does not exist in endemic equilibrium as shown in Figure 8, which means that the disease has been eliminated.
(ⅳ) When perturbation parameters (θ1,θ2) fall on the curve SN, system (2.1) has a unique endemic equilibrium which is a saddle node. When the parameters cross through curve SN from region Ⅰ to Ⅱ or Ⅳ, the system undergoes a saddle-node bifurcation, and the number of endemic equilibrium changes from zero to two, which are shown in Figures 9 and 10, respectively.
(ⅴ) When perturbation parameters (θ1,θ2) cross curve H into area Ⅲ, the system (2.1) undergoes a supercritical Hopf bifurcation of the endemic equilibrium point E2 and produces an unstable limit cycle as displayed in Figure 11. If the initial value of (x,y) is within the limit cycle rather than in an endemic equilibrium E2, the number of infectious individuals periodically varies.
(ⅵ) When the parameters (θ1,θ2) change from left to right in region Ⅲ, the limit cycle becomes bigger and larger, then reaches equilibrium E2 and becomes a homoclinic loop at the moment when (θ1,θ2) lies on the curve HL. After the parameter (θ1,θ2) is across the curve HL, the homoclinic loop breaks up. Figure 12 illustrates this phenomenon.
In this paper, the purpose of our research is to explore the influence of psychological reactions of people on the spread of disease. We choose the SIS model to describe the spread mechanism of some infectious diseases that do not give survivors immunity. In the model, the component that can reflect the influence of psychological behavior on the spread of disease is the incidence function. Many researchers have studied various incidence functions in SIS models.
Based on pioneering research work, in 2019, a generalized nonmonotone and saturated incidence function kI21+βI+αI2 in the SIRS(Susceptible-Infectious-Recovered-Infectious) system was proposed by Lu et al. [7]; they considered that the incidence function, which describes the infection force, should not be just monotonic, nonmonotonic, or saturated, but a combination of monotonicity, nonmonotonicity, and saturation properties. Taking into account the psychological and crowding effect, the incidence function seemed reasonable to describe the infection force of some specific infectious diseases. However, we have found that the basic reproduction number of the SIRS model with the incidence rate kI21+βI+αI2 is zero, but the disease can still be persistent; therefore, how to calculate the basic reproduction number of this model is still an open question. Therefore, the incidence rate is difficult to understand. Furthermore, the biological meaning of the parameter β is not defined, and the condition β>−2√α is mandatory.
So, we propose a more reasonable incidence rate k1I+k2I21+aI2S with the combination of monotonicity, nonmonotonicity, and saturation properties. When k1=0, the incidence function f(I)=k1I+k2I21+aI2 becomes the saturated incidence function in [22], which increases monotonously and then increases to k2a as I→∞. When k1>0, f(I) increases first and then decreases to k2a as I→∞, which describes the fact that the infection force of some infectious diseases grows rapidly to maximum as a new disease emerges or an old disease reemerges, and then trends to a value.
In this paper, we conducted a qualitative analysis. The basic reproduction number R0 of the model (1.1) is bk1d(d+ε+μ), and we present that model (1.1) can undergo backward bifurcation if we take R0 as the bifurcation parameter if Λ>q−qw. Backward bifurcation was proposed by Castillo-Chavez and Song [23] to illustrate that even if the basic reproduction number is R0<1, disease outbreaks are still possible. The backward bifurcation has further epidemiological implications by providing a threshold Ra; when Λ>q−qw, model (1.1) shows bistable behavior (endemic equilibrium E2 and disease-free equilibrium are stable) if Ra<R0<1, and the model has a unique endemic equilibrium E2 and the disease-free equilibrium becomes unstable if R0>1. Moreover, a saddle-node bifurcation at the threshold Ra has been studied. In our paper, there is the basic reproduction number R0, the threshold Ra, and the critical values k∗1 and k10 of k1, which measure the linear infection risk to determine the dynamic of the system (2.1).
Briefly, when R0≤1 and k1k2, which measures the proportion of linear over nonlinear infection hazards, is greater than or equal to bd+ε, the disease can be eliminated for all initial populations. When R0>1 or R0=1 and k1k2<bd+ε, the disease will persist in the form of multiple periodic coexistent oscillations bifurcated from the equilibrium E2, or coexisting steady states for some initial populations. When k1k2<bd+ε, system (2.1) will present complex dynamics, including backward, saddle-node, Hopf bifurcation, etc. When k∗1<k1<k10, the disease will disappear if R0<Ra, but the disease will persist if R0>Ra, and for R0=Ra, these conditions are not enough to determine dynamical behaviors, which implies the disease will persist or die out, which depends on the values of independent parameters k1 and k2, and requires further evaluation.
Moreover, we have proved that the model undergoes Hopf bifurcation and undergoes Bogdanov-Takens bifurcation of, at most, codimension two, which indicates that the disease may exhibit multiple homeostasis and periodic outbreaks. However, our discussion of the nature of equilibrium E2 is not exhaustive and needs further study. If the first Lyapunov coefficient l1 is equal to zero, the equilibrium E2 may be a weak focus with a multiplicity of at least two; meanwhile, if the parameters are suitable, there may be two limit cycles in system (2.1). Due to the complex computation, we will not discuss it here.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This work was supported by the National Natural Science Foundation of China (12171135), the Research Funding for High-Level Innovative Talents of Hebei University (801260201242); and the Innovation Capacity Enhancement Program-Science and Technology Platform Project, Hebei Province (22567623H).
The authors declare there is no conflicts of interest.
[1] |
M. De la Sen, R. Nistal, S. Alonso-Quesada, A. Ibeas, Some formal results on positivity, stability, and endemic steady-state attainability based on linear algebraic tools for a class of epidemic models with eventual incommensurate delays, Discrete Dyn. Nat. Soc., 2019 (2019), 8959681. https://doi.org/10.1155/2019/8959681 doi: 10.1155/2019/8959681
![]() |
[2] |
M. De la Sen, A. Ibeas, S. Alonso-Quesada, R. Nistal, On a new epidemic model with asymptomatic and dead-infective subpopulations with feedback controls useful for ebola disease, Discrete Dyn. Nat. Soc., 2017 (2017), 4232971. https://doi.org/10.1155/2017/4232971 doi: 10.1155/2017/4232971
![]() |
[3] |
W. O. Kermack, A. G. McKendrick, Contributions to the mathematical theory of epidemics. Ⅱ. — The problem of endemicity, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci., 138 (1932), 55–83. https://doi.org/10.1098/rspa.1932.0171 doi: 10.1098/rspa.1932.0171
![]() |
[4] |
V. Capasso, G. Serio, A generalization of the Kermack-Mckendrick deterministic epidemic model, Math. Biosci., 42 (1978), 43–61. https://doi.org/10.1016/0025-5564(78)90006-8 doi: 10.1016/0025-5564(78)90006-8
![]() |
[5] | D. M. Xiao, Y. G. Zhou, Qualitative analysis of an epidemic model, Can. Appl. Math. Q., 14 (2006), 469–492. |
[6] |
Y. G. Zhou, D. M. Xiao, Y. L. Li, Bifurcations of an epidemic model with non-monotonic incidence rate of saturated mass action, Chaos, Solitons Fractals, 32 (2007), 1903–1915. https://doi.org/10.1016/j.chaos.2006.01.002 doi: 10.1016/j.chaos.2006.01.002
![]() |
[7] |
M. Lu, J. C. Huang, S. G. Ruan, P. Yu, Bifurcation analysis of an SIRS epidemic model with a generalized nonmonotone and saturated incidence rate, J. Differ. Equations, 267 (2019), 1859–1898. https://doi.org/10.1016/j.jde.2019.03.005 doi: 10.1016/j.jde.2019.03.005
![]() |
[8] |
P. van den Driessche, J. Watmough, A simple SIS epidemic model with a backward bifurcation, J. Math. Biol., 40 (2000), 525–540. https://doi.org/10.1007/s002850000032 doi: 10.1007/s002850000032
![]() |
[9] |
M. H. Alharbi, Global investigation for an "SIS" model for COVID-19 epidemic with asymptomatic infection, Math. Biosci. Eng., 20 (2023), 5298–5315. https://doi.org/10.3934/mbe.2023245 doi: 10.3934/mbe.2023245
![]() |
[10] |
Z. Shuai, P van den Driessche, Global stability of infectious disease models using Lyapunov functions, SIAM J. Appl. Math., 73 (2013), 1513–1532. https://doi.org/10.1137/120876642 doi: 10.1137/120876642
![]() |
[11] |
P van den Driessche, J. Watmough, Reproduction numbers and sub-threshold endemic equilibria for compartmental models of disease transmission, Math. Biosci., 180 (2002), 29–48. https://doi.org/10.1016/S0025-5564(02)00108-6 doi: 10.1016/S0025-5564(02)00108-6
![]() |
[12] | Y. A. Kuznetsov, Numerical Analysis of Bifurcations, Cambridge University Press, 2004. |
[13] | Z. Zhang, Qualitative Theory of Differential Equations, Science Press, 1992. |
[14] | S. Wiggins, Introduction to Applied Nonlinear Dynamical Systems and Chaos, 2nd edition, Springer, 2003. |
[15] |
J. C. Huang, Y. J. Gong, S. G. Ruan, Bifurcation analysis in a predator-prey model with constant-yield predator harvesting, Discrete Contin. Dyn. Syst. Ser. B, 18 (2013), 2101–2121. https://doi.org/10.3934/dcdsb.2013.18.2101 doi: 10.3934/dcdsb.2013.18.2101
![]() |
[16] |
Y. L. Tang, D. Q. Huang, S. G. Ruan, W. N. Zhang, Coexistence of limit cycles and homoclinic loops in a SIRS model with a nonlinear incidence rate, SIAM J. Appl. Math., 69 (2008), 621–639. https://doi.org/10.1137/070700966 doi: 10.1137/070700966
![]() |
[17] | B. D. Hassard, N. D. Kazarinoff, Y. H. Wan, Theory and Applications of Hopf bifurcation, Cambridge University Press, 1981. |
[18] | J. E. Marsden, M. McCracken, The Hopf Bifurcation and its Applications, Springer-Verlag, 2012. |
[19] |
R. I. Bogdanov, Versal deformations of a singular point of a vector field on the plane in the case of zero eigenvalues, Funct. Anal. Appl., 9 (1975), 144–145. https://doi.org/10.1007/BF01075453 doi: 10.1007/BF01075453
![]() |
[20] | F. Takens, Forced oscillations and bifurcations, in Global Analysis of Dynamical Systems, CRC Press, (2001), 11–71. |
[21] | L. Perko, Differential Equations and Dynamical Systems, Springer Science & Business Media, 2013. |
[22] |
S. G. Ruan, W. D. Wang, Dynamical behavior of an epidemic model with a nonlinear incidence rate, J. Differ. Equations, 188 (2003), 135–163. https://doi.org/10.1016/S0022-0396(02)00089-X doi: 10.1016/S0022-0396(02)00089-X
![]() |
[23] |
C. Castillo-Chavez, B. J. Song, Dynamical models of tuberculosis and their applications, Math. Biosci. Eng., 1 (2004), 361–404. https://doi.org/10.3934/mbe.2004.1.361 doi: 10.3934/mbe.2004.1.361
![]() |